Mitochondrial ion channels as oncological targets

0 downloads 0 Views 2MB Size Report
The HK-VDAC complex (more specifically, HK) can also be targeted by the alkylating agent 3-bromopyruvate (3BP) that has been reported efficacious in in vitro ...
Mitochondrial ion channels as oncological targets L Leanza 1 , M Zoratti 2 , E Gulbins 3 and I Szabo 1 M itochondria, the key bioenergetic intracellular organelles, harbor a number of proteins with proven or hypothetical ion channel functions. Growing evidence points to the important contribution of these channels to the regulation of mitochondrial function, such as ion homeostasis imbalances profoundly affecting energy transducing processes, reactive oxygen species production and mitochondrial integrity. Given the central role of mitochondria in apoptosis, their ion channels with the potential to compromise mitochondrial function have become promising targets for the treatment of malignancies. Importantly, in vivo evidence demonstrates the involvement of the proton-transporting uncoupling protein, a mitochondrial potassium channel, the outer membrane located porin and the permeability transition pore in tumor progression/control. In this review, we focus on mitochondrial channels that have been assigned a definite role in cell death regulation and possess clear oncological relevance. Overall, based on in vivo and in vitro genetic and pharmacological evidence, mitochondrial ion channels are emerging as promising targets for cancer treatment. Keywords: mitochondria; ion channels; direct targeting; pharmacology M ITOCHONDRIA; ION HOM EOSTASIS AND APOPTOSIS; IM PACTS ON SIGNALING AND CANCER The mitochondrial inner membrane (IM M ) surrounding the matrix permits the formation of an electrochemical proton gradient that drives aerobic adenosine triphosphate (ATP) synthesis, and the outer membrane (OM M ) delimits an intermembrane space (IM S) between the OM M and IM M . Proteins with fundamental roles in cell death, such as cytochrome c and SM AC/Diablo, can be released from the IM S upon receipt of a proapoptotic signal and can lead to commitment of the cell to apoptosis. The released serine protease HtrA2/Omi, apoptosis-inducing factor and endo- nuclease G also contribute to the executive phase of apoptosis. 1 M uch effort has been devoted to identifying drugs capable of inducing cancer cell death by stimulating OM M rupture/ permeabilization and release of these proteins. 2 M itochondrial ion channels are therapeutically relevant to cancer because they can influence both OM M and IM M permeability. M embers of the B-cell lymphoma 2 (Bcl-2) family and M AC (mitochondrial apoptosis-induced channel) are directly involved in OM M permeabilization, but they do not function as bona fide ion channels. Excellent reviews are available on their role in this process; 3,4 therefore, the present review focuses on the mitochondrial voltage-dependent anion channel (VDAC, also known as porin) among the OM M channels. Selective induction of IM M permeabilization in cancer cells is also a potential therapeutic strategy because mitochondrial energy production becomes compromised when the IM M becomes freely permeable to solutes, which leads to cell death. As a consequence of changes in ion flux across the IM M , alterations of mitochondrial function occur that produce severe effects on the overall fitness, including ATP production, of these organelles. M ost available data on ion channels or ion flux pathways that modulate cancer cells can be ultimately reconciled with the idea that the perturbation of ion homeostasis within mitochondria leads to cancer cell death via reactive oxygen species (ROS) production, mitochondrial calcium overload and metabolism changes, but other mechanisms of action might also exist. Indeed, mitochondria are crucial for maintaining intracellular Ca 2 þ homeostasis and produce ROS, which are two important factors for cancer cell survival and proliferation. A detailed overview of the above processes is beyond the scope of this review; thus, we briefly describe the general mechanisms by which ion channels may affect ROS production, mitochondrial calcium homeostasis and metabolism. The matrix-negative difference in electrical potential across the IM M (DC m , ranging between À 150 and À 180 mV) is maintained by the proton pumps of the respiratory chain (RC). Consequently, cations (K þ and Ca 2 þ ) flow from the IM S (where the concentra- tion of ions is comparable to cytosolic concentrations because the OM M is permeable to these ions) to the matrix when a permeation pathway opens, as was demonstrated several years ago by experiments with ionophores such as valinomycin. The electrical- equivalent anion outflow is considered much less relevant because the anion pool in the mitochondrial matrix is limited compared with the cytosolic K þ pool. To compensate for charge movement (unless permeating counterions are also present, which is feasible in experimental settings but unusual under physiological conditions), the RC must increase the rate of proton transfer from the matrix to the IM S. To increase RC activity according to the chemiosmotic model, the transmembrane electrochemical proton gradient (D~ m H , composed of mainly DC m ) must decrease. Thus, passive charge flow and D~ m H (DC m ) are coupled, and the opening of K þ or Ca 2 þ channels in the IM M will lead to depolarization. Clearly, this model also functions in reverse; closing cation channels is expected (and observed) to lead to an increase of a D~ m H (DC m ) (that is, hyperpolarization). 1

Department of Biology, University of P adova, P adova, Italy; 2 CNR Institute of Neuroscience and Department of Biomedical Sciences, University of P adova, P adova, Italy and

Department of Molecular Biology, University of Duisburg-Essen, Essen, Germany. Correspondence: P rofessor I Szabo, Department of Biology, University of P adova, Viale G Colombo 3, P adova P D, 35138, Italy. E-mail: [email protected] or [email protected] 3

Such alterations may in turn have consequences on the rate of superoxide formation. 5,6 Hyperpolarization lowers the efficacy of cytochrome c oxidase, 7 induces the reduction of RC complexes and intermediates and thereby increases the probability of a one- electron transfer to oxygen at respiratory complex I. Complex-III- dependent ROS formation can also be significant at high respiration rates (depolarized conditions) because of a high concentration of semiquinone at center ‘O’ (SQ o or Q P ) of the bc 1 complex. 8 Thus, enhanced superoxide (O 2 À ) production occurs in both the matrix and IM S. 7,9–11 In addition, recent data have highlighted that complex II also generates superoxide or hydrogen peroxide molecules at the flavin site II F when electrons are supplied by either succinate or the reduced ubiquinone pool. 12 Irrespective of the site of production, which might depend on the substrates being oxidized, 13 the superoxide anion is the precursor of most ROS. ROS can diffuse to different subcellular compartments. Increased ROS production is important for the maintenance and evolution of the cancerous phenotype. 14 The effects of ROS vary according to the stage of carcinogenesis. In the tumor initiation phase, ROS may cause DNA damage and mutagenesis, whereas in established cancer they may act as signal mediators through the Akt pathway and stimulate proliferation, conferring a growth advantage. 15 ROS may also inhibit Pyruvate Kinase M 2, the embryonic form produced by cancer cells, thereby diverting reducing equivalents to the pentose phosphate pathway. 16 Conversely, ROS release can induce cell death by damaging proteins, lipids and DNA and thus act as anticancer agents. 17–19 In addition, the activation of pro-apoptotic kinases, such as apoptosis signal-regulating kinase 1 (ASK1) and death-associated protein kinases (DAPK; 5 members), 20,21 by ROS is of fundamental importance. 15,17–21 Oxidative stress can also activate the pro- apoptotic Bax and permeability transition pore (M PTP) (see below), 22 whose prolonged opening leads to the loss of the mitochondrial D~ m H , mitochondrial swelling, cristae remodeling and pro-apoptotic factor release. 23 Bax and other Bcl-2 family proteins regulate OM M permeabilization through pore formation and possible regulation of mitochondrial fission and fusion. 24 In light of this dichotomy, both antioxidant and pro-oxidant approaches have been proposed to antagonize cancer. 25–28 Antioxidants are expected to be useful for cancer prevention, whereas pro-oxidants are desirable to potentially and selectively induce cancer cell death. Chronic metabolic oxidative stress, likely occurring in most cancer cell types, can indeed be exploited because these cells exhibit an increased sensitivity to ROS-induced apoptosis. 26 Therefore, therapies that specifically target the RC, either directly or indirectly, to further elevate ROS production should selectively kill cancer cells. Because pharmacological modulation of mitochondrial ion channel function is expected to lead to changes in membrane potential, targeting these channels to increase ROS levels above a critical threshold might be one way to induce cancer cell apoptosis, as detailed below. Furthermore, altered ion fluxes might lead, via hyperpolariza- tion, to an increased driving force for Ca 2 þ entry into the matrix, whereas cation (K þ or M g 2 þ ) channel opening-induced depolar- ization decreases Ca 2 þ uptake. M itochondria have the capacity to accumulate Ca 2 þ , using the mitochondrial calcium uniporter as a highly selective ion channel, 29 to a much higher concentration than found in the cytoplasm. Thus, mitochondria shape cytosolic calcium signals. 30 The mitochondrial matrix possesses a Ca 2 þ buffering system and calcium exit pathways (the Na þ /Ca 2 þ exchanger and transient opening of the permeability transition pore) that normally prevents excessive [Ca 2 þ ] buildup (see, for example, Drago et al. 31 ). However, after apoptotic stimuli, sustained Ca 2 þ release from the endoplasmic reticulum (‘ER stress’) may lead to high mitochondrial matrix Ca 2 þ levels that causes prolonged opening of the permeability transition pore and subsequent loss of mitochondrial D~ m H . Thus, Ca 2 þ overload disrupts energy production, releases proapoptotic proteins into the cytoplasm and enables apoptosis and/or necrosis. 31–33 In addition, mitochondrial Ca 2 þ accumulation also leads to enhanced ROS production through multiple potential mechanisms. 34 Although the exact mechanistic role of mitochondrial Ca 2 þ homeostasis in tumorigenesis (beyond ROS production and modulation of apoptosis) is not well understood, mitochondrial Ca 2 þ is an important signal transducer that can initiate pathways mediating cell death. Therefore, drugs increasing calcium uptake may efficiently kill cancer cells; however, how to selectively target cancer cells instead of healthy cells by inducing mitochondrial calcium accumulation remains unclear. M itochondrial calcium uptake is important for oxidative phosphorylation regulation 35 and energy metabolism control because it enhances the rate of NADH production by modulating the enzymes of the Szent-Gyo ¨ rgyi–Krebs cycle and fatty acid oxidation pathways. 36,37 In general, strategies that increase energy and metabolite production in energy-demanding cancer cells promote a high glycolytic flux rate and thus allow enhanced tumor cell growth. 38,39 Some IM M channels, by regulating DC, matrix volume, matrix acidification and ROS production, would be expected to impact the functionality of mitochondrial RC complexes. Indeed, enhanced substrate oxidation with matrix volume expansion has been demonstrated, 40 and pharmacological activation of calciuminduced potassium channels by NS11021, which affects mitochondrial matrix volume, leads to enhanced respiratory control. 41 Interestingly, the respiration rate and oxidative phosphorylation efficiency were found to be increased in August rats that are characterized by an increased rate of ATPdependent mitochondrial potassium transport as compared with Wistar rats that have lower ATP-dependent potassium channel activity. 42 A recent paper reported the physical interaction and functional coupling of the mitochondrial large- conductance calcium-activated potassium channel (mtBK Ca ) and complex IV of the RC. 43 Whether this physical interaction also occurs in other channels and plays a direct role in the regulation of mitochondrial energy fitness by ion channels is an important and unresolved question. Finally, the changes in matrix volume resulting from channel opening are expected to lead to structural cristae reorganization that has recently been shown to affect respiratory supercomplex organization and respiration efficiency. 44 Along with IM M channels, a role for the OM M channel VDAC in metabolism regulation has become clear; the glycolytic enzyme hexokinase (HK), when bound to VDAC1, regulates metabolite trafficking (including ATP) 45,46 through the OM M channel and provides cancer cells with metabolic advantages. Furthermore, VDAC1-bound HK is less sensitive to inhibition by its product, glucose-6-phosphate, 47 and avoids product inhibition that further boosts a high glycolytic rate. Thus, IM M and OM M channels are expected to affect mitochondrial metabolism and RC function through multiple mechanisms. Therefore, their modulation may lead to cancer cell metabolism alteration, ultimately promoting cell death, as discussed below. In summary, OM M and IM M channel modulation may affect cancer cell fate through different signaling mechanisms. THE IN VITRO AND IN VIVO TARGETING OF M ITOCHONDRIAL ION CHANNELS IN CANCER The mitochondrial channels characterized over the past two decades include the outer membrane-localized VDAC and, in the IM M , potassium channels such as mtK ATP , mtBK Ca , mtIK Ca , mtKv1.3, mtTASK-3, the nonselective permeability transition pore M PTP, chloride channels, the magnesium-permeable M rs-2, the calcium uniporter, uncoupling proteins and proteins that have been shown to function as ion channels in other membranes but not

2

yet directly observed in mitochondrial membranes. 48–51 Here, we discuss the in vitro and in vivo evidence for ion channels involved in cancer cell death obtained through either pharmacological or genetic means (Figure 1). Data from animal models clearly demonstrate the importance of mitochondrial channels and their potential as therapeutic cancer targets. The possible roles of individual channels in cell death in relation to the processes discussed above are summarized in Figure 2 and are described in greater detail below. Pharmacological targeting of these channels (summarized in Table 1) and possible pitfalls are also critically discussed. The outer membrane-localized VDAC The major protein of the OM M , VDAC or porin, is required for cancer cell survival and participates in apoptotic signaling. A large body of literature examines the role of VDAC in the regulation of apoptosis. 48,52 VDAC is being studied as a cancer-specific target because tumor cells have increased glycolysis and VDAC expression. 53 The role of VDAC1 and the other isoforms, VDAC2 and VDAC3, in cell death is complex, 54–56 but importantly, in vivo evidence shows that in cancer cells, the association of HK with VDAC1 protects against mitochondrial-mediated apoptosis. Therefore, disruption of the HKVDAC1 complex represents an attractive therapeutic cancer target. Overexpression of hexokinase-1 and 2 (HK2) and their association with VDAC are important characteristics of glycolytic cancers. 57 The expression of VDACs has also been found to be elevated in cancerous cells compared with normal cells and may be altered with chemotherapy. High VDAC levels are an unfavorable prognostic factor; 53 moreover, VDAC downregulation by RNA interference reduces cancer growth. 58 This evidence may seem at odds with the finding that VDAC overexpression induces apoptosis, 59– 61 but it illustrates how the functional meaning of a biological parameter depends on context. VDAC upregulation in cancer goes hand in hand with HK2 upregulation and can be considered a component of glycolytic upregulation. HK2 binding to VDAC,

which allows for ATP transport out of mitochondria, leads to a cancer cell metabolic advantage (termed the Warburg effect), and it antagonizes cell death through the inhibition of Bax-induced cytochrome c release 62,63 and/or inhibition of the mitochondrial permeability transition (M PT). 64 HK dissociation from VDAC seems to favor cell

3

4

death through the disruption of aerobic glycolysis and the cell’s energy balance by altering the interaction of Bcl-2 family proteins with mitochondria and by regulation of ROS production and facilitation of VDAC oligomer formation. 48,55 A class of novel apoptosis-inducing anticancer drugs, which currently comprises small molecules such as clotrimazole, bifonazole (both fungicides, in the tens-of-micromolar range) and methyl jasmonate (M J), act by disrupting the HK-VDAC1 complex (Table 1). Among these small molecules, M J has been demonstrated to be active in preclinical models of melanoma, prostate and breast cancer, lymphoblastic leukemia and multiple myeloma, 65,66 also independently of p53 status and has been tested in at least one pilot human study. 67 However, the major obstacle for adopting jasmonates as anticancer agents is their relatively high dosage needed to exert their action. It should be also noted that different mechanisms of action from the mechanism described above have also been envisioned. M J activates Caspase-3 through ROS production, 68 and another study showed that it also downregulates the expression of proliferating cell nuclear antigen. 69 Furthermore, M J can bind to members of the Aldoketo reductase family 1, 70 which are potential tumor biomarkers involved in steroid metabolism. 71,72 The fact that M J does not affect nontransformed, intact cells is compatible with most suggested mechanisms. Whether M J can also induce cell death in the absence of the proapoptotic Bax/Bak, which occurs in chemoresistant cancer cells, remains unknown. Clotrimazole could have multiple effects because it also acts as a membrane- permeable inhibitor of the KCa3.1 calcium-dependent potassium channel that has also been localized to the IM M 73 and seems to be involved in apoptosis regulation. 74 In addition, peptides corresponding to the amino-terminus of both HK-I 75 and HK-II 62 and a cell-permeable HK-II-based peptide 76 may be potentially useful as pharmacological tools. The HK-VDAC complex (more specifically, HK) can also be targeted by the alkylating agent 3-bromopyruvate (3BP) that has been reported efficacious in in vitro and in vivo studies (for example, see reviews in refs 77–80) and possesses strong potential as an anticancer agent in humans. 79,81 In animal models, 3BP showed high efficacy against advanced-stage malignant tumors by inhibiting both glycolysis and mitochondrial energy generation, possibly by interfering with the HK-VDAC1 complex. 80 Thus, 3BP has been proposed to target metabolism and block energy supplies. In addition, GAPDH, 82 components of the RC, 83 ER and lysosomes are targets of this compound. Its remarkable specificity for cancer cells may be partially attributed to the presence of glycolytic pathway (HK and GAPDH) members among these targets; however, most of 3BP specificity likely stems from facilitated entry into glycolytic cancer cells via the lactate transporter. 84 The transcription factor ATF2, which elicits oncogenic and tumor-suppressor activities in

5

melanoma and nonmalignant skin cancer, respectively, has been recently identified as a potent disruptor of the HK1–VDAC1 complex. When localized to mitochondria in a protein kinase C-E- dependent manner, ATF2 increases mitochondrial permeability and promotes apoptosis. 85 The heterocycle erastin has been found to selectively induce the death of cells with mutations in the oncogenes HRas, KRas and BRaf in a VDAC2/3-dependent manner. Apparently, this compound induces oxidative stress involving VDAC through an unknown mechanism. 86 An erastin homolog, PRLX 93936, is currently being evaluated in a clinical trial in multiple myeloma patients (http://www.cancer.gov/clinicaltrials). Finally, the anticancer agent furanonaphthoquinone, isolated from Avicennia plants, induces caspase-dependent apoptosis via ROS production by acting on VDAC. VDAC has also been proposed to mediate exit of superoxide anions from the intermembrane space, and its closure caused internal oxidative stress and sensitized mito- chondria to Ca 2 þ /ROSinduced M PT. 87 M oreover, the anticancer activity of furanonaphthoquinone and ROS production were enhanced by VDAC1 overexpression. 88 VDAC has also been proposed to directly participate in OM M permeabilization and mediate cytochrome c release, either by oligomer formation 48 or by the formation of a large pore comprising VDAC and Bax/Bak (Tsujimoto et al. 89 but see M artinez-Caballero et al. 90 ). Knockdown of VDAC1 prevented cisplatin-induced conformational activation of Bax 91 or selenite- induced cytochrome c release. 92 However, Bax-induced cytochrome c release from mitochondria isolated from wild-type or VDAC1-, VDAC3- and VDAC1/VDAC3-null cells was reported to be identical. 93 In any case, the binding of anti-apoptotic Bcl-2 and Bcl-xL to VDAC1 (with resulting porin activity inhibition) 94 produced anti-apoptotic actions. 95 A recent study reported that constructs consisting of cell-penetrating Antp fused to VDAC1- derived sequences, which prevented the interaction of VDAC1 and HK, Bcl-2 and BclxL, were effective in the selective eradication of B cells from patients with chronic lymphocytic leukemia. 96 In summary, a systematic search for compounds that act on VDAC to antagonize cancer remains to be performed, but those compounds already identified are promising. Inner membrane-localized permeability transition pore A number of cytotoxic agents and cellular stressors trigger the loss of IM M permeability. This process is most often caused by M PT that is considered to be a final common pathway of various forms of cell death. 97,98 In fact, M PT results in a ‘bioenergetic catastrophe’; the transmembrane electrochemical proton gradient dissipates, ATP synthesis ceases and respiratory substrates are lost from the mitochondrial matrix. M PT is caused by the opening of a large, Ca 2 þ -activated and oxidative stress-sensitive pore (M PTP) that creates an IM M permeable to ions and solutes up to 1500 Da molecular weight and leads to matrix swelling. This pore coincides with the mitochondrial megachannel that has been studied by patch clamp and characterized by conductances up to 1.5 nS. 99 M PTP has recently been proposed to be formed by dimers of the FoF1 ATP synthase and Cyclophilin D. 98,100 M PTP/mitochondrial megachannel opening is considered to account for a substantial portion of the tissue damage caused by ischemia/reperfusion and oxidative stress. In cancer cells, signaling pathways are activated that render mitochondria more resistant to M PT induction. 101–103 Because chemotherapeutic agents cause oxidative stress, they may activate signals that induce cell death through M PTP opening. 104 M PT inducers have potential oncological applications, but potential side effects involving the nervous system or cardiac tissues can be expected. A consistent number of compounds (often used at relatively high concentrations) have been shown to induce oxidative stress and/or disruption of Ca 2 þ homeostasis along with M PT. Some of the M PTP-targeting molecules, such as 4-(N-(S-glutathionylacetyl) amino) phenylarsenoxide, are currently being evaluated in clinical trials as promising drugs against refractory tumors. 105,106 M itochondria-penetrating peptides, such as mastoparan-like sequences, peptides of the innate immune system and molecules developed by the Kelley group, 107–109 also induce M PT and could become candidates for future clinical trials. M PTP can also be indirectly activated (or inhibited) through modulatory signaling pathways. For example, in cancer models containing constitutively active extracellular signal-regulated kinase, this kinase acts through the glycogen synthase kinase- 3b/Cyclophilin D axis to repress cell death by M PT inducers such as arachidonic acid or BH3-mimetic EM 20-25. 110 Induction of oxidative stress by the gold complex AUL12 can lead to activation of glycogen synthase kinase-3a/b that favors M PTP opening. 111 A large portion of M PTP-opening inducers (direct or indirect) for which in vivo antitumor activities have been reported 2 are natural compounds, including jasmonates, 112–114 betulinic acid, the synthetic retinoid CD437, 115,116 berberine, 117,118 honokiol, 119,120 a-bisabolol 121 and shikonin 122 (Table 1). Among these com- pounds, betulinic acid is currently in a phase I /II clinical trial for dysplastic nevi (http://www.cancer.gov/clinicaltrials). A retinoid analog, NRX 195183, is also in a phase II trial for acute promyelocytic leukemia (http://www.cancer.gov/clinicaltrials). A specific, powerful M PTP inhibitor is also available: cyclosporin A, a cyclic endecapeptide. 123–125 Cyclosporin A inhibits the M PTP through binding to matrix cyclophilin, a peptidyl-prolyl cis–trans isomerase, and acts as an immunosuppressant by inhibiting calcineurin. Further evidence for the link between M PTP and cancer is illustrated by the observation that patients treated with cyclosporin A to prevent rejection of organ transplants have a high incidence of cancer. 126 The IM M potassium channel Kv1.3 As mentioned above, IM M -localized potassium channels are predicted to participate in the regulation of mitochondrial membrane potential, volume and ROS production. Our research groups have identified a potassium channel, mtKv1.3, in the IM M of several cell types and have shown with mitochondria from cells expressing or lacking this channel that mtKv1.3 activity indeed has an impact on DC m and on ROS production. 127,128 Similar to some other IM M channels, 50 mtKv1.3 is the mitochondrial counterpart of the plasma membrane-localized Shaker family potassium channel, Kv1.3. Its crucial role in apoptosis became evident because the expression of a mitochondria-targeted Kv1.3 construct was sufficient to induce cell death upon apoptotic stimuli in CTLL-2 T lymphocytes that lack Kv channels and are otherwise resistant to apoptosis. A physical interaction between Bax and mtKv1.3 has been demonstrated in apoptotic cells, and Bax has been shown to inhibit channel activity at nM concentrations. 129,130 Incubation of Kv1.3-positive isolated mitochondria with Bax or specific mtKv1.3 inhibitors triggered apoptotic events, including membrane potential changes (hyperpolarization followed by depolarization because of M PTP opening), ROS production and cytochrome c release; however, Kv1.3-deficient mitochondria were resistant to these inhibitors. The highly conserved Bax lysine residue 128 was shown to protrude into the intermembrane space. 131 This residue is responsible for the inhibitory effect of Bax on Kv1.3 activity because it mimics a crucial lysine residue in Kv1.3blocking peptide toxins. Indeed, mutant BaxK128E did not exert its effects on Kv1.3 and mitochondria, and it did not mediate apoptosis in Bax/Bak double-knockout mouse embryonic fibroblasts, indicating that the toxin-like action of Bax on Kv1.3 triggers the above described mitochondrial events. 130 Three membrane-permeable inhibitors of Kv1.3, Psora-4, PAP-1 and clofazimine, are able to induce cell death at mM concentra- tions by directly targeting mtKv1.3 in different cancer cell lines, in

6

contrast to the membrane-impermeable high-affinity Kv1.3 inhibitors ShK and margatoxin. 132,133 Genetic deficiency or small interfering RNA-mediated downregulation of Kv1.3 abrogated the effects of the membrane-permeable drugs that also killed cells in the absence of Bax and Bak, in agreement with the above described model (Figure 3). In vivo, intraperitoneal injection of clofazimine in an orthotopic melanoma B16F10 mouse model reduced tumor size by 90%, without causing obvious side effects. 133 Recently, promising results were obtained with primary human cancer cells from patients with chronic lymphocytic leukemia, where drugs significantly and exclusively decreased the survival of pathologic B cells (which express higher levels of Kv1.3 compared with B cells from normal subjects) by inducing the crossing of a critical ROS threshold. 134 Thus, the selective apoptosis-inducing action of these drugs on tumor cells seems to be related to a synergistic effect of an altered cancer cell redox state and higher Kv1.3 expression. 134,135 Interestingly, As 2 O 3 , a clinically active antileukemic agent, has been reported to inhibit mitochondrial respiratory function, increase ROS generation and enhance the activity of other O 2 . À -producing agents against cultured and primary patient leukemia cells. 136 Because clofazimine is already used clinically to treat some autoimmune diseases and leprosy 137 and shows an excellent safety profile, targeting mtKv1.3 is a feasible cancer therapy for at least some cancer types. Interestingly, altered Kv1.3 expression is observed in several different cancer cell lines and tumors, 138 and a correlation between Kv1.3 expression and cell sensitivity to clofazimine has been reported. 139 In addition, the ability of these drugs to induce cell death in the absence of Bax and Bak, similar to other agents, 140–142 could be useful because downregulation of these pro-apoptotic proteins represents a common tumor cell resistance mechanism. 143–146 Calcium-dependent potassium channels The activity of an intermediate conductance potassium channel (IK mitochondrial

Ca

, KCa3.1) has been recorded from the inner

membranes of human cancer cells. 73,147 . IK Ca can be selectively inhibited by low concentrations of clotrimazole and TRAM -34. This latter drug has been reported to synergistically increase the sensitivity of melanoma cells to the death receptor ligand TRAIL (tumor necrosis factor-related apoptosis-inducing ligand) via its action on mtIK Ca . 74 However, the effects of TRAM -34 on mitochondrial bioenergetics have not yet been investigated in detail, but TRAM -34 has been reported to induce hyper- polarization of the mitochondrial membrane (which is expected if an influx of positively charged ions/molecules is inhibited), confirming that functional IK Ca is expressed in the IM M . Interestingly, TRAM -34 application induced Bax translocation to the mitochondria, representing an early step of apoptosis. Both TRAIL and TRAM -34 are characterized by a relatively good safety profile, suggesting that co-administration of these two drugs might be exploited to treat melanoma. In addition to the intermediate conductance channels, the presence of small-conductance calciumactivated potassium (SK2/ KCa2.2) channels in the IM M has been recently reported. 148,149 Its pharmacological activation in a neuronal cell line exposed to toxic glutamate levels attenuated the loss of the mitochondrial transmembrane potential, blocked mitochondrial fission, prevented the release of proapoptotic mitochondrial proteins and reduced cell death. 149 SK2 channel opening prevented mitochondrial calcium overload and mitochondrial superoxide formation. It will be interesting to investigate whether the application of mtSK2 inhibitors alone or in combination with other chemotherapeutics could promote cell death in cancer cells expressing SK2. To our knowledge, whether the mitochondrial large-conduc- tance Ca 2 þ -activated K þ channel (BK Ca /KCa1.1) 150 has a role in cancer has not been investigated. However, the channel opener CGS7184 has been reported to induce the glioma cell death that was accompanied by increased respiration and mitochondrial depolarization, 151 perhaps downstream of an increased Ca 2 þ release from the ER. 152

7

TWIK-related acid-sensitive K þ channel-3 (TASK-3) A protein recognized by an antibody against TASK-3 (TWIK-related acid-sensitive K þ channel-3; KCNK9), a two-pore potassium channel, was identified in the mitochondria of melanoma, keratinocytes 153 and healthy intestinal epithelial cells. 154 The reduction of TASK-3 expression in WM 35 melanoma cells compromised mitochondrial function and cell survival. 155 TASK-3 knockdown also decreased human keratinocyte HaCaT cell viability following ultraviolet B irradiation. 156 In mitochondria isolated from HaCaT cells, TASK-3-compatible activity has been recorded for the first time by electrophysiology (patch clamp). 156 Two aspects of TASK-3 still require investigation: first, whether the observed migration and invasion-reducing effects of TASK-3 overexpression in breast cancer cells 157 and increased apoptosis induced by TASK-3 blockers (zinc and methanandamide) in ovarian carcinoma 158 are correlated with the mitochondrial channel expression and, second, whether pharmacological action on mtTASK-3 can affect cancer cell death. However, because no highly specific modulators are available, this task remains difficult. Uncoupling protein (UCP) Uncoupling proteins (UCP1–5) belong to the superfamily of mitochondrial anion-carrier proteins, and UCP1 has recently been shown to mediate fatty acid-regulated proton transport. 159 The widely expressed UCP-2 has also been proposed to transport protons, 160 (for review, see Cannon and Nedergaard 161 ), although its physiological importance and the occurrence of UCP2- mediated proton leak are not entirely clear and remain debatable. UCP2 has been suggested to regulate cell survival by decreasing mitochondrial ROS because it might limit the maximal value of the proton gradient across the IM M , and a depolarizing proton leak would be expected to diminish superoxide production 162,163 (see, for example, Cannon et al. 164 for a discussion on this topic). Increased ROS production was observed in UCP2 knockout mice, whereas UCP2 overexpression may contribute to a higher apoptotic threshold promoting cancer cell survival because it prevents oxidative injury. Indeed, UCP2 overexpression, documented in numerous tumor types, including leukemia, ovarian, bladder, esophageal, testicular, colorectal, kidney, pancreatic, lung and prostate tumors, has been shown to protect cells from oxidative stress 165 and even to abolish the apoptosis-inducing effects of chemotherapeutic drugs. 166 Interestingly, UCP2 overexpression has also been proposed to directly contribute to the Warburg phenotype 167 and to the development of tumors in an orthotopic in vivo model of breast cancer. 168 In support of this hypothesis, UCP2 expression in M CF7 breast cancer cells was shown to lead to a decreased mitochondrial membrane potential and increased tumorigenic properties. These studies suggest that UCP2 overexpression is involved in the development of a variety of cancers and that UCP2 can be considered as a promising oncological target, although the actual ions transported by UCP2 under physiological conditions are still uncertain. As recently suggested, 162 UCP2 may act as an uniporter for pyruvate, presumably promoting pyruvate efflux from the matrix and thus restricting glucose availability for mitochondrial respiration 162 and promoting mitochondrial fatty acid oxidation. Efflux of pyruvate would aid highly glycolytic cells, in which large amounts of pyruvate would otherwise put enormous redox pressure on the mitochondria. Nevertheless, a plant-derived small molecule, genipin, has been identified as an agent capable of suppressing the tumor-promoting functions of UCP2 presumably by abolishing UCP2-mediated proton leak. 168 In sharp contrast with the above studies, when highly metastatic M DA-M B-231 breast cancer cells recombinantly overexpressing UCP1, UCP2 or UCP3 were injected into the flanks of athymic nude mice, in vivo tumor growth was reduced 2.2-fold compared with empty-vector control cells. 169 The authors interpreted their results as a consequence of UCP-induced increased autophagy and decreased mitochondrial function. This unexpected observation warrants further investigation. M itochondrial calcium uniporter M CU The mitochondrial Ca 2 þ ‘uniporter’ (M CU), a calcium-selective ion channel, 29 is responsible for low-affinity calcium uptake into the mitochondrial matrix. The recent identification of the protein responsible for this activity 170,171 has produced new perspectives pertaining to the control of Ca 2 þ signaling in the cell. In fact, M CU may be a very useful tool to influence a number of cellular calcium-dependent processes, including cell death. 51 For example, subthreshold apoptotic stimulation in synergy with cytosolic Ca 2 þ waves induced M PTP opening 172 and, moreover, M CU overexpression resulted in increased apoptosis upon with H 2 O 2 and C2-ceramide challenges. 170 In addition, recent data suggest that in the absence of M ICU1, a proposed negative regulator of M CU, mitochondria become constitutively loaded with Ca 2 þ , resulting in excessive ROS generation and sensitivity to apoptotic stress. 173 To avoid calcium overload, M CU activity is also regulated by a dominant-negative subunit, M CUb. 174 As for a possible relationship between M CU and cancer, the M CU-targeting microRNA miR-25 has been shown to affect Ca 2 þ homeostasis in colon cancer cells through the specific downregulation of M CU. The miR-25 caused a strong decrease in mitochondrial Ca 2 þ uptake and, importantly, conferred resistance to Ca 2 þ -dependent apoptotic challenges. 175 Thus, M CU could be an important protein for tumorigenesis, and its specific pharmacological activators, if identified, might become useful tools. However, the situation is likely more complex. M CU overexpression has been observed in breast cancer samples and has been suggested to provide a survival advantage against some cell death pathways. 176 Pharmacological or small interfering RNA-mediated inhibition of mitochondrial Ca 2 þ entry sensitized cancer cells to positively charged gold nanoparticles through the induction of endoplasmic reticulum stress. 177 Further complications arose from recent work reporting that epidermal growth-induced epithelial– mesenchymal transition in breast cancer cells was not associated with changes in M CU expression, as assessed by quantitative reverse transcriptase–PCR. 178 Finally, in sharp contrast with the extensive literature suggesting that alterations in mitochondrial calcium play a central role in acute metabolic regulation and determination of a threshold for cell death, the basal metabolism of knockout animals was not markedly altered in the absence of M CU. 179 Furthermore, although experiments with M CU À / À mitochondria confirmed that M CU is required for calcium- induced M PTP opening, M CU À / À mice exhibited no evidence of protection from ischemia–reperfusion-mediated injury. In this study, tumor incidence in knockout mice was not assessed. The forthcoming elucidation of M CU regulation, 180–182 as well as the creation of a conditional knockout mouse, might provide explanation for these apparently contrasting findings. Acid-sensing ion channel 1a In a recent study, the sodium-permeable acid-sensing ion channel 1a (ASIC1a) was documented in mitochondria, although the transport properties of this mitochondrial protein were not defined. 183 Interestingly, neurons from ASIC1a À / À mice were resistant to cytochrome c release and inner mitochondrial membrane depolarization, and mitochondria from these cells displayed an enhanced Ca 2 þ retention capacity and oxidative stress response. The authors proposed that mitochondrial ASIC1a may serve as a regulator of M PTP through a still ill-defined mechanism, thus contributing to oxidative neuronal cell death. Interestingly, lack of ASIC1 prevented axonal degeneration in experimental autoimmune encephalomyelitis; however, M PTP

8

THERAPEUTIC PERSPECTIVES The concept that mitochondria can serve as promising pharma- cological targets in oncology is gaining increasing support. One intervention strategy envisions direct targeting of mitochondrial channels to induce the death of unwanted, cancerous cells. Although much work remains in this relatively new field, studies involving VDAC, inner membrane K þ and Ca 2 þ channels and the permeability transition pore justify high expectations. To produce advances in this field, further progress should focus on under- standing the functions of these and other channels in normal cells, mechanisms involved in cell death induction upon their manip- ulation, characterization of other mitochondrial channels (for example, the 107-pS anion channel) in an oncological context, screening of potential drugs and improvements specifically targeting drugs to mitochondria and to cancerous cells. Table 1 summarizes the currently available pharmacological data target- ing mitochondrial ion channels obtained in various cancer cells. To expand this field, studies should address the expression level of a set of ion channels in cancer tissues. For example, among the channels that are also present in the mitochondrial membranes, IK Ca (KCa3.1) expression has been shown to be significantly increased in breast tumors bearing a p53 mutation, and a correlation was detected between high expression of this channel and high tumor grades. 185 The same situation occurs for VDAC1, VDAC2 and VDAC3. 185 Conversely, BK Ca (KCa1.1) expression was reduced in patients with higher grade tumors compared with patients with lower grade tumors. 185 Emerging data on mitochondrial ion channels might help to identify specific drugs that can be used to obtain maximal cancer-killing efficacy. The major hurdle from a pharmacological perspective concerns the specificity and membrane permeability of drugs acting on mitochondrial channels. In many cases, the used inhibitors have multiple off-target effects as well (for example, methyl jasmonate, honokiol). Because several channels are present in both the plasma membrane and the mitochondrial membrane, drugs targeting the mitochondrial channels might exert undesired effects on processes linked to plasma membrane channels of healthy cells (for example, proliferation). Furthermore, only cell membrane-permeable drugs will reach mitochondrial channels, and a portion of these drugs is likely to be retained in cellular membranes. Therefore, higher drug concentrations than those concentrations sufficient to block the channels are needed to exert an effect on cell viability, and these concentrations may lead to a loss of specificity. Different strategies can be envisioned to create more efficient targeting of mitochondrial channels. Cell- penetrating peptides (such as the VDAC-related peptides) are one possibility, but clinical problems could arise, as these peptides are able to cross the blood–brain barrier. Another strategy might be derivatization of the drugs by adding a permeating positively charged moiety 186–188 that is able to drive accumulation of the drugs in mitochondria. In summary, to exploit mitochondrial channels as oncological targets, we should explore exactly how these ion channels contribute to cell fate, and state-of-the-art pharmacological tools should be developed and used for specific targeting. CONFLICT OF INTEREST The authors declare no conflict of interest.

ACKNOWLEDGEM ENTS We are grateful to all co-authors for their important contributions to the work carried out in their laboratories and reported in this review. We apologize for not citing all publications that have been published on this topic. The work carried out in the authors’ laboratories was supported in part by grants from the Italian Association for Cancer Research (AIRC; Grant 11814 to IS), the EMBO Young Investigator P rogram grant (to IS), the P rogetti di Rilevante Interesse Nazionale (P RIN) program (2010CSJX4F to IS and 20107Z8XBW_004 to MZ), the Fondazione Cassa di Risparmio di P adova e Rovigo (to MZ), the CNR P roject of Special Interest on Aging (to MZ), the DFG Grant Gu 335/13–3 (to EG), the International Association for Cancer Research (to EG) and the P rogetto Giovani Studiosi 2012 (to LL).

REFERENCES 1 Galluzzi L, Kepp O, Kroemer G. Mitochondria: master regulators of danger signalling. Nat Rev Mol Cell Biol 2012; 13: 780–788. 2 Fulda S, Galluzzi L, Kroemer G. Targeting mitochondria for cancer therapy. Nat Rev Drug Discov 2010; 9: 447–464. 3 Bender T, Martinou JC. Where killers meet--permeabilization of the outer mito- chondrial membrane during apoptosis. Cold Spring Harb P erspect Biol 2013; 5: a011106. 4 Dejean LM, Ryu SY, Martinez-Caballero S, Teijido O, P eixoto P M, KW Kinnally. MAC and Bcl-2 family proteins conspire in a deadly plot. Biochim Biophys Acta 2010; 1797: 1231–1238. 5 Starkov AA. The role of mitochondria in reactive oxygen species metabolism and signaling. Ann NY Acad Sci 2008; 1147: 37–52. 6 Murphy MP . How mitochondria produce reactive oxygen species. Biochem J 2009; 417: 1–13. 7 Kadenbach B. Intrinsic and extrinsic uncoupling of oxidative phosphorylation. Biochim Biophys Acta 2003; 1604: 77–94. 8 Malinska D, Mirandola SR, Kunz WS. Mitochondrial potassium channels and reactive oxygen species. FEBS Lett 2010; 584: 2043–2048. 9 Kowaltowski AJ, de Souza-P into NC, Castilho RF, Vercesi AE. Mitochondria and reactive oxygen species. Free Radic Biol Med 2009; 47: 333–343. 10 Demin OV, Kholodenko BN, Skulachev VP . A model of O 2 .-generation in the complex III of the electron transport chain. Mol Cell Biochem 1998; 184: 21–33. 11 O-Uchi J, Ryu SY, Jhun BS, Hurst S, Sheu SS. Mitochondrial ion channels/trans- porters as sensors and regulators of cellular redox signaling. Antioxid Redox Signal 2013; P MID: 24180309. 12 Orr AL, Quinlan CL, P erevoshchikova IV, Brand MD. A refined analysis of super- oxide production by mitochondrial sn-glycerol 3phosphate dehydrogenase. J Biol Chem 2012; 287: 42921–42935. 13 Quinlan CL, P erevoshchikova IV, Hey-Mogensen M, Orr AL, Brand MD. Sites of reactive oxygen species generation by mitochondria oxidizing different sub- strates. Redox Biol 2013; 1: 304–312. 14 Fiaschi T, Chiarugi P . Oxidative stress, tumor microenvironment, and metabolic reprogramming: a diabolic liaison. Int J Cell Biol 2012; 2012: 762825. 15 Biasutto L, Szabo I, Zoratti M. Mitochondrial effects of plant-made compounds. Antioxid Redox Signal 2012; 15: 3039–3059. 16 Anastasiou D, P oulogiannis G, Asara JM, Boxer MB, Jiang JK, Shen M et al. Inhibition of pyruvate kinase M2 by reactive oxygen species contributes to cellular antioxidant responses. Science 2011; 334: 1278–1283. 17 Ueda S, Masutani H, Nakamura H, Tanaka T, Ueno M, Yodoi J. Redox control of cell death. Antioxid Redox Signal 2002; 4: 405–414. 18 Circu ML, Aw TY. Reactive oxygen species, cellular redox systems, and apoptosis. Free Radic Biol Med 2010; 48: 749–762. 19 P anieri E, Gogvadze V, Norberg E, Venkatesh R, Orrenius S, Zhivotovsky B. Reactive oxygen species generated in different compartments induce cell death, survival, or senescence. Free Radic Biol Med 2013; 57: 176–187. 20 Bialik S, Kimchi A. The death-associated protein kinases: structure, function, and beyond. Annu Rev Biochem 2006; 75: 189–210. 21 Takeda K, Noguchi T, Naguro I, Ichijo H. Apoptosis signal-regulating kinase 1 in stress and immune response. Annu Rev P harmacol Toxicol 2008; 48: 199–225. 22 Ghibelli L, Diederich M. Multistep and multitask Bax activation. Mitochondrion 2010; 10: 604–613. 23 Scorrano L. Opening the doors to cytochrome c: changes in mitochondrial shape and apoptosis. Int J Biochem Cell Biol 2009; 41: 1875–1883. 24 Landes T, Martinou JC. Mitochondrial outer membrane permeabilization during apoptosis: the role of mitochondrial fission. Biochim Biophys Acta 2011; 1813: 540–545. 25 Trachootham D, Alexandre J, Huang P . Targeting cancer cells by ROS-mediated mechanisms: a radical therapeutic approach? Nat Rev Drug Discov 2009; 8: 579–591. 26 Ralph SJ, Rodrı ´guez-Enrı ´quez S, Neuzil J, Saavedra E, Moreno-Sa ´nchez R. The causes of cancer revisited: ‘ mitochondrial malignancy’ and ROS-induced onco- genic transformation - why mitochondria are targets for cancer therapy. Mol Aspects Med 2010; 31: 145–170.

9

27 Ralph SJ, Moreno-Sa ´nchez R, Neuzil J, Rodrı ´guez-Enrı ´quez S. Inhibitors of succinate: quinone reductase/Complex II regulate production of mitochondrial reactive oxygen species and protect normal cells from ischemic damage but induce specific cancer cell death. P harm Res 2011; 28: 2695–2730. 28 Wang J, Yi J. Cancer cell killing via ROS: to increase or decrease, that is the question. Cancer Biol Ther 2008; 7: 1875–1884. 29 Kirichok Y, Krapivinsky G, Clapham DE. The mitochondrial calcium uniporter is a highly selective ion channel. Nature 2004; 427: 360–364. 30 Gunter TE, Buntinas L, Sparagna G, Eliseev R, Gunter K. Mitochondrial calcium transport: mechanisms and functions. Cell Calcium 2000; 28: 285–296. 31 Drago I, P izzo P , P ozzan T. After half a century mitochondrial calcium in- and efflux machineries reveal themselves. EMBO J 2011; 30: 4119–4125. 32 Kinnally KW, P eixoto P M, Ryu SY, Dejean LM. Is mMP TP the gatekeeper for necrosis, apoptosis, or both? Biochim Biophys Acta 2011; 1813: 616–622. 33 Rasola A, Bernardi P . Mitochondrial permeability transition in Ca( 2 þ )-dependent apoptosis and necrosis. Cell Calcium 2011; 50: 222–233. 34 P eng TI, Jou MJ. Oxidative stress caused by mitochondrial calcium overload. Ann NY Acad Sci 2010; 1201: 183–188. 35 Maack C, Cortassa S, Aon MA, Ganesan AN, Liu T, O’ Rourke B. Elevated cytosolic Na þ decreases mitochondrial Ca 2 þ uptake during excitation-contraction cou- pling and impairs energetic adaptation in cardiac myocytes. Circ Res 2006; 99: 172–182. 36 Nichols BJ, Denton RM. Towards the molecular basis for the regulation of mitochondrial dehydrogenases by calcium ions. Mol Cell Biochem 1995; 149-150: 203–212. 37 Denton RM. Regulation of mitochondrial dehydrogenases by calcium ions. Bio- chim Biophys Acta 2009; 1787: 1309–1316. 38 P edersen P L. The cancer cell’ s ‘ power plants’ as promising therapeutic targets: an overview. J Bioenerg Biomembr 2007; 39: 1–12. 39 Wallace DC. Mitochondria and cancer. Nat Rev Cancer. 2012; 12: 685–698. 40 Halestrap AP . The regulation of the oxidation of fatty acids and other substrates in rat heart mitochondria by changes in the matrix volume induced by osmotic strength, valinomycin and Ca2 þ . Biochem J 1987; 244: 159–164. 41 Aon MA, Cortassa S, Wei AC, Grunnet M, O’ Rourke B. Energetic performance is improved by specific activation of K þ fluxes through K(Ca) channels in heart mitochondria. Biochim Biophys Acta 2010; 1797: 71–80. 42 Venediktova N, Shigaeva M, Belova S, Belosludtsev K, Belosludtseva N, Gorbacheva O, Lezhnev E, Lukyanova L, Mironova G. Oxidative phosphorylation and ion transport in the mitochondria of two strains of rats varying in their resistance to stress and hypoxia. Mol Cell Biochem 2013; 383: 261–269. 43 Bednarczyk P , Wieckowski MR, Broszkiewicz M, Skowronek K, Siemen D, Szewczyk A. P utative structural and functional coupling of the mitochondrial BKCa channel to the respiratory chain. P LoS One 2013; 8: e68125. 44 Cogliati S, Frezza C, Soriano ME, Varanita T, Quintana-Cabrera R, Corrado M et al. Mitochondrial cristae shape determines respiratory chain supercomplexes assembly and respiratory efficiency. Cell 2013; 155: 160–171. 45 P erevoshchikova IV, Zorov SD, Kotova EA, Zorov DB, Antonenko YN. Hexokinase inhibits flux of fluorescently labeled ATP through mitochondrial outer mem- brane porin. FEBS Lett 2010; 584: 2397–2402. 46 Lemasters JJ, Holmuhamedov E. Voltage-dependent anion channel (VDAC) as mitochondrial governator--thinking outside the box. Biochim Biophys Acta 2006; 1762: 181–190. 47 Azoulay-Zohar H, Israelson A, Abu-Hamad S, Shoshan-Barmatz V. In self-defence: hexokinase promotes voltage-dependent anion channel closure and prevents mitochondriamediated apoptotic cell death. Biochem J 2004; 377: 347–355. 48 Shoshan-Barmatz V, De P into V, Zweckstetter M, Raviv Z, Keinan N, Arbel N. VDAC, a multi-functional mitochondrial protein regulating cell life and death. Mol Aspects Med 2010; 31: 227–285. 49 Zoratti M, De Marchi U, Gulbins E, Szabo ` I. Novel channels of the inner mito- chondrial membrane. Biochim Biophys Acta 2009; 1787: 351–363. 50 Szabo ` I, Leanza L, Gulbins E, Zoratti M. P hysiology of potassium channels in the inner membrane of mitochondria. P flugers Arch 2012; 463: 231–246. 51 Rizzuto R, De Stefani D, Raffaello A, Mammucari C. Mitochondria as sensors and regulators of calcium signalling. Nat Rev Mol Cell Biol 2012; 13: 566–578. 52 Tan W. VDAC blockage by phosphorothioate oligonucleotides and its implication in apoptosis. Biochim Biophys Acta 2012; 1818: 1555–1561. 53 Grills C, Jithesh P V, Blayney J, Zhang SD, Fennell DA. Gene expression meta-analysis identifies VDAC1 as a predictor of poor outcome in early stage non-small cell lung cancer. P LoS One 2011; 6: e14635. 54 McCommis KS, Baines CP . The role of VDAC in cell death: friend or foe? Biochim Biophys Acta, 2012; 1818: 1444–1450. 55 Shoshan-Barmatz V, Golan M. Mitochondrial VDAC1: function in cell life and death and a target for cancer therapy. Curr Med Chem 2012; 19: 714–735. 56 Shoshan-Barmatz V, Mizrachi D. VDAC1: from structure to cancer therapy. Front Oncol 2012; 2: 164. 57 Wolf A, Agnihotri S, Micallef J, Mukherjee J, Sabha N, Cairns R et al. Hexokinase 2 is a key mediator of aerobic glycolysis and promotes tumor growth in human glioblastoma multiforme. J Exp Med 2011; 208: 313–326. 58 Koren I, Raviv Z, Shoshan-Barmatz V. Downregulation of voltage-dependent anion channel-1 expression by RNA interference prevents cancer cell growth in vivo. Cancer Biol Ther 2010; 9: 1046–1052. 59 Zaid H, Abu-Hamad S, Israelson A, Nathan I, Shoshan-Barmatz V. The voltage- dependent anion channel-1 modulates apoptotic cell death. Cell Death Differ 2005; 12: 751–760. 60 Mader A, Abu-Hamad S, Arbel N, Gutie ´rrez-Aguilar M, Shoshan-Barmatz V. Dominant-negative VDAC1 mutants reveal oligomeric VDAC1 to be the active unit in mitochondria-mediated apoptosis. Biochem J 2010; 429: 147–155. 61 Sharaf el dein O, Gallerne C, Brenner C, Lemaire C. Increased expression of VDAC1 sensitizes carcinoma cells to apoptosis induced by DNA cross-linking agents. Biochem P harmacol 2012; 83: 1172–1182. 62 P astorino JG, Shulga N, Hoek JB. Mitochondrial binding of hexokinase II inhibits Bax-induced cytochrome c release and apoptosis. J Biol Chem 2002; 277: 7610–7618. 63 Gall JM, Wong V, P imental DR, Havasi A, Wang Z, P astorino JG et al. Hexokinase regulates Bax-mediated mitochondrial membrane injury following ischemic stress. Kidney Int 2011; 79: 1207–1216. 64 Chiara F, Castellaro D, Marin O, P etronilli V, Brusilow WS, Juhaszova M et al. Hexokinase II detachment from mitochondria triggers apoptosis through the permeability transition pore independent of voltage-dependent anion channels. P LoS One 2008; 3: e1852. 65 Fingrut O, Flescher E. P lant stress hormones suppress the proliferation and induce apoptosis in human cancer cells. Leukemia 2002; 16: 608–616. 66 Klippel S, Jakubikova J, Delmore J, Ooi M, McMillin D, Kastritis E et al. Methyl- jasmonate displays in vitro and in vivo activity against multiple myeloma cells. Br J Haematol 2012; 159: 340–351. 67 P almieri B, Iannitti T, Capone S, Flescher E. A preliminary study of the local treatment of preneoplastic and malignant skin lesions using methyl jasmonate. Eur Rev Med P harmacol Sci 2011; 15: 333–336. 68 Kim JH, Lee SY, Oh SY, Han SI, P ark HG, Yoo MA et al. Methyl jasmonate induces apoptosis through induction of Bax/Bcl-XS and activation of caspase-3 via ROS production in A549 cells. Oncol Rep 2004; 12: 1233–1238. 69 Tong QS, Jiang GS, Zheng LD, Tang ST, Cai JB, Liu Y et al. Methyl jasmonate downregulates expression of proliferating cell nuclear antigen and induces apoptosis in human neuroblastoma cell lines. Anticancer Drugs 2008; 19: 573–581. 70 Davies NJ, Hayden RE, Simpson P J, Birtwistle J, Mayer K, Ride JP et al. AKR1C isoforms represent a novel cellular target for jasmonates alongside their mitochondrial-mediated effects. Cancer Res 2009; 69: 4769–4775. 71 Ji Q, Chang L, VanDenBerg D, Stanczyk FZ, Stolz A. Selective reduction of AKR1C2 in prostate cancer and its role in DHT metabolism. P rostate 2003; 54: 275–289. 72 Vihko P , Herrala A, Ha ¨ rko ¨ nen P , Isomaa V, Kaija H, Kurkela R et al. Enzymes as modulators in malignant transformation. J Steroid Biochem Mol Biol 2005; 93: 277–283. 73 De Marchi U, Sassi N, Fioretti B, Catacuzzeno L, Cereghetti GM, Szabo ` I et al. Intermediate conductance Ca 2 þ -activated potassium channel (KCa3.1) in the inner mitochondrial membrane of human colon cancer cells. Cell Calcium 2009; 45: 509–516. 74 Quast SA, Berger A, Buttsta ¨ dt N, Friebel K, Scho ¨ nherr R, Eberle J. General sen- sitization of melanoma cells for TRAIL-induced apoptosis by the potassium channel inhibitor TRAM-34 depends on release of SMAC. P LoS One 2012; 7: e39290. 75 Gelb BD, Adams V, Jones SN, Griffin LD, MacGregor GR, McCabe ER. Targeting of hexokinase 1 to liver and hepatoma mitochondria. P roc Natl Acad Sci USA 1992; 89: 202–206. 76 Majewski N, Nogueira V, Bhaskar P , Coy P E, Skeen JE, Gottlob K et al. Hexokinase- mitochondria interaction mediated by Akt is required to inhibit apoptosis in the presence or absence of Bax and Bak. Mol Cell 2004; 16: 819–830. 77 P edersen P L. 3-Bromopyruvate (3BP ) a fast acting, promising, powerful, specific, and effective ‘ small molecule’ anti-cancer agent taken from labside to bedside: introduction to a special issue. J Bioenerg Biomembr 2012; 44: 1–6. 78 Shoshan MC. 3-Bromopyruvate: targets and outcomes. J Bioenerg Biomembr 2012; 44: 7–15. 79 Ko YH, Verhoeven HA, Lee MJ, Corbin DJ, Vogl TJ, P edersen P L. A translational study ‘ ‘ case report’ ’ on the small molecule ‘ ‘ energy blocker’ ’ 3-bromopyruvate (3BP ) as a potent anticancer agent: from bench side to bedside. J Bioenerg Biomembr 2012; 44: 163–170. 80 Cardaci S, Desideri E, Ciriolo MR. Targeting aerobic glycolysis: 3-bromopyruvate as a promising anticancer drug. J Bioenerg Biomembr 2012; 44: 17–29. 81 Galluzzi L, Kepp O, Heiden MG, Kroemer G. Metabolic targets for cancer therapy. Nat Rev Drug Discov 2013; 12: 829–846.

10

82 Tang Z, Yuan S, Hu Y, Zhang H, Wu W, Zeng Z et al. Over-expression of GAP DH in human colorectal carcinoma as a preferred target of 3-bromopyruvate propyl ester. J Bioenerg Biomembr 2012; 44: 117–125. 83 Rodrigues-Ferreira C, da Silva AP , Galina A. Effect of the antitumoral alkylating agent 3-bromopyruvate on mitochondrial respiration: role of mitochondrially bound hexokinase. J Bioenerg Biomembr 2012; 44: 39–49. 84 Ko YH, Smith BL, Wang Y, P omper MG, Rini DA, Torbenson MS et al. Advanced cancers: eradication in all cases using 3-bromopyruvate therapy to deplete ATP . Biochem Biophys Res Commun 2004; 324: 269–275. 85 Lau E, Kluger H, Varsano T, Lee KY, Scheffler I, Rimm DL et al. P KCE promotes oncogenic functions of ATF2 in the nucleus while blocking its apoptotic function at mitochondria. Cell 2012; 148: 543–555. 86 Yagoda N, von Rechenberg M, Zaganjor E, Bauer AJ, Yang WS, Fridman DJ et al. RAS-RAF-MEK-dependent oxidative cell death involving voltage-dependent anion channels. Nature 2007; 447: 864–868. 87 Tikunov A, Johnson CB, P ediaditakis P , Markevich N, Macdonald JM, Lemasters JJ et al. Closure of VDAC causes oxidative stress and accelerates the Ca( 2 þ )- induced mitochondrial permeability transition in rat liver mitochondria. Arch Biochem Biophys 2010; 495: 174–181. 88 Simamura E, Shimada H, Hatta T, Hirai K. Mitochondrial voltage-dependent anion channels (VDACs) as novel pharmacological targets for anti-cancer agents. J Bioenerg Biomembr 2008; 40: 213–217. 89 Tsujimoto Y, Shimizu S. VDAC regulation by the Bcl-2 family of proteins. Cell Death Differ 2000; 7: 1174–1181. 90 Martinez-Caballero S, Dejean LM, Kinnally MS, Oh KJ, Mannella CA, Kinnally KW. Assembly of the mitochondrial apoptosis-induced channel, MAC. J Biol Chem 2009; 284: 12235–12245. 91 Tajeddine N, Galluzzi L, Kepp O, Hangen E, Morselli E, Senovilla L et al. Hierarchical involvement of Bak, VDAC1 and Bax in cisplatin-induced cell death. Oncogene 2008; 27: 4221–4232. 92 Tomasello F, Messina A, Lartigue L, Schembri L, Medina C, Reina S et al. Outer membrane VDAC1 controls permeability transition of the inner mito- chondrial membrane in cellulo during stress-induced apoptosis. Cell Res 2009; 19: 1363–1376. 93 Baines CP , Kaiser RA, Sheik T, Craigen WJ, Molkentin JD. Voltage-dependent anion channels are dispensable for mitochondrial-dependent cell death. Nat Cell Biol 2007; 9: 550–555. 94 Shimizu S, Ide T, Yanagida T, Tsujimoto Y. Electrophysiological study of a novel large pore formed by Bax and the voltage-dependent anion channel that is permeable to cytochrome c. J Biol Chem 2000; 275: 12321–12325. 95 Arbel N, Ben-Hail D, Shoshan-Barmatz V. Mediation of the antiapoptotic activity of Bcl-xL protein upon interaction with VDAC1 protein. J Biol Chem 2012; 287: 23152–23161. 96 P rezma T, Shteinfer A, Admoni L, Raviv Z, Sela I, Levi I, Shoshan-Barmatz V. VDAC1-based peptides: novel pro-apoptotic agents and potential therapeutics for B-cell chronic lymphocytic leukemia. Cell Death Dis 2013; 4: e809. 97 Brenner C, Grimm S. The permeability transition pore complex in cancer cell death. Oncogene 2006; 25: 4744–4756. 98 Bernardi P . The mitochondrial permeability transition pore: a mystery solved? Front P hysiol 2013; 4: 95. 99 Szabo ´ I, Zoratti M. The giant channel of the inner mitochondrial membrane is inhibited by cyclosporin A. J Biol Chem 1991; 266: 3376–3379. 100 Giorgio V, von Stockum S, Antoniel M, Fabbro A, Fogolari F, Forte M et al. Dimers of mitochondrial ATP synthase form the permeability transition pore. P roc Natl Acad Sci USA 2013; 110: 5887–5892. 101 Rasola A, Sciacovelli M, P antic B, Bernard P . Signal transduction to the perme- ability transition pore. FEBS Lett 2010; 584: 1989–1896. 102 Traba J, Del Arco A, Duchen MR, Szabadkai G, Satru ´ stegui J. SCaMC-1 promotes cancer cell survival by desensitizing mitochondrial permeability transition via ATP /ADP -mediated matrix Ca( 2 þ ) buffering. Cell Death Differ 2012; 19: 650–660. 103 Matassa DS, Amoroso MR, Maddalena F, Landriscina M, Esposito F. New insights into TRAP 1 pathway. Am J Cancer Res 2012; 2: 235–248. 104 Chiara F, Gambalunga A, Sciacovelli M, Nicolli A, Ronconi L, Fregona D et al. Chemotherapeutic induction of mitochondrial oxidative stress activates GSK-3a/ b and Bax, leading to permeability transition pore opening and tumor cell death. Cell Death Dis 2012; 3: e444. 105 Elliott MA, Ford SJ, P rasad E, Dick LJ, Farmer H, Hogg P J et al. P harmaceutical development of the novel arsenical based cancer therapeutic GSAO for P hase I clinical trial. Int J P harm 2012; 426: 67–75. 106 Brenner C, Moulin M. P hysiological roles of the permeability transition pore. Circ Res 2012; 111: 1237– 1247. 107 Jones S, Martel C, Belzacq-Casagrande AS, Brenner C, Howl J. Mitoparan and target-selective chimeric analogues: membrane translocation and intracellular redistribution induces mitochondrial apoptosis. Biochim Biophys Acta 2008; 1783: 849–863. 108 Risso A, Braidot E, Sordano MC, Vianello A, Macrı ` F, Skerlavaj B et al. BMAP -28, an antibiotic peptide of innate immunity, induces cell death through opening of the mitochondrial permeability transition pore. Mol Cell Biol 2002; 22: 1926–1935. 109 Horton KL, P ereira MP , Stewart KM, Fonseca SB, Kelley SO. Tuning the activity of mitochondria-penetrating peptides for delivery or disruption. Chembiochem 2012; 13: 476–485. 110 Rasola A, Sciacovelli M, Chiara F, P antic B, Brusilow WS, Bernardi P . Activation of mitochondrial ERK protects cancer cells from death through inhibition of the permeability transition. P roc Natl Acad Sci USA 2010; 107: 726–731. 111 Chiara F, Rasola A. GSK-3 and mitochondria in cancer cells. Front Oncol 2013; 3: 16. 112 Flescher E. Jasmonates--a new family of anti-cancer agents. Anticancer Drugs 2005; 16: 911–916. 113 Flescher E. Jasmonates in cancer therapy. Cancer Lett 2007; 245: 1–10. 114 Raviv Z, Cohen S, Reischer-P elech D. The anti-cancer activities of jasmonates. Cancer Chemother P harmacol 2013; 71: 275–285. 115 Lena A, Rechichi M, Salvetti A, Bartoli B, Vecchio D, Scarcelli V et al. Drugs targeting the mitochondrial pore act as cytotoxic and cytostatic agents in temozolomide-resistant glioma cells. J Transl Med 2009; 7: 13. 116 Javadov S, Hunter JC, Barreto-Torres G, P arodiRullan R. Targeting the mito- chondrial permeability transition: cardiac ischemia-reperfusion versus carcino- genesis. Cell P hysiol Biochem 2011; 27: 179–190. 117 P ereira GC, Branco AF, Matos JA, P ereira SL, P arke D, P erkins EL et al. Mitochondrially targeted effects of berberine [Natural Yellow 18, 5,6-dihydro- 9,10-dimethoxybenzo(g)-1,3-benzodioxolo(5,6-a) quinolizinium] on K1735-M2 mouse melanoma cells: comparison with direct effects on isolated mitochondrial fractions. J P harmacol Exp Ther 2007; 323: 636–649. 118 P ereira CV, Machado NG, Oliveira P J. Mechanisms of berberine (natural yellow 18)-induced mitochondrial dysfunction: interaction with the adenine nucleotide translocator. Toxicol Sci 2008; 105: 408–417. 119 Li L, Han W, Gu Y, Qiu S, Lu Q, Jin J et al. Honokiol induces a necrotic cell death through the mitochondrial permeability transition pore. Cancer Res 2007; 67: 4894–4903. 120 Dong JX, Zhao GY, Yu QL, Li R, Yuan L, Chen J et al. Mitochondrial dysfunction induced by honokiol. J Membr Biol 2013; 246: 375–381. 121 Cavalieri E, Bergamini C, Mariotto S, Leoni S, P erbellini L, Darra E et al. Involvement of mitochondrial permeability transition pore opening in alpha-bisabolol induced apoptosis. FEBS J 2009; 276: 3990–4000. 122 Han W, Li L, Qiu S, Lu Q, P an Q, Gu Y et al. Shikonin circumvents cancer drug resistance by induction of a necroptotic death. Mol Cancer Ther 2007; 6: 1641–1649. 123 Broekemeier KM, Dempsey ME, P feiffer DR. Cyclosporin A is a potent inhibitor of the inner membrane permeability transition in liver mitochondria. J Biol Chem 1989; 264: 7826–7830. 124 Crompton M, Ellinger H, Costi A. Inhibition by cyclosporin A of a Ca 2 þ -depen- dent pore in heart mitochondria activated by inorganic phosphate and oxidative stress. Biochem. J 1988; 255: 357–360. 125 Fournier N, Ducet G, Crevat A. Action of cyclosporine on mitochondrial calcium fluxes. J Bioenerg Biomembr 1987; 19: 297–303. 126 Norman KG, Canter JA, Shi M, Milne GL, Morrow JD, Sligh JE. Cyclosporine A suppresses keratinocyte cell death through MMP TP inhibition in a model for skin cancer in organ transplant recipients. Mitochondrion 2010; 10: 94–101. 127 Szabo ` I, Bock J, Jekle A, Soddemann M, Adams C, Lang F et al. A novel potassium channel in lymphocyte mitochondria. J Biol Chem 2005; 280: 12790–12798. 128 Gulbins E, Sassi N, Grassme ` H, Zoratti M, Szabo ` I. Role of Kv1.3 mitochondrial potassium channel in apoptotic signalling in lymphocytes. Biochim Biophys Acta 2010; 1797: 1251–1259. 129 Szabo ´ I, Bock J, Grassme ´ H, Soddemann M, Wilker B, Lang F et al. Mitochondrial potassium channel Kv1.3 mediates Bax-induced apoptosis in lymphocytes. P roc Natl Acad Sci USA 2008; 105: 14861–14866. 130 Szabo ` I, Soddemann M, Leanza L, Zoratti M, Gulbins E. Single-point mutations of a lysine residue change function of Bax and Bcl-xL expressed in Bax- and Bak- less mouse embryonic fibroblasts: novel insights into the molecular mechanisms of Bax-induced apoptosis. Cell Death Differ 2011; 18: 427–438. 131 Annis MG, Soucie EL, Dlugosz P J, Cruz-Aguado JA, P enn LZ, Leber B et al. Bax forms multispanning monomers that oligomerize to permeabilize membranes during apoptosis. EMBO J 24: 2096–2103. 132 Leanza L, Zoratti M, Gulbins E, Szabo ` I. Induction of apoptosis in macro- phages via Kv1.3 and Kv1.5 potassium channels. Curr Med Chem 2012; 19: 5394–5404. 133 Leanza L, Henry B, Sassi N, Zoratti M, Chandy KG, Gulbins E et al. Inhibitors of mitochondrial Kv1.3 channels induce Bax/Bak-independent death of cancer cells. EMBO Mol Med 2012; 4: 577–593.

11

134 Leanza L, Trentin L, Becker KA, Frezzato F, Zoratti M, Semenzato G et al. Clofazimine, P sora-4 and P AP -1, inhibitors of the potassium channel Kv1.3, as a new and selective therapeutic strategy in chronic lymphocytic leukemia. Leukemia 2013; 27: 1782–1785. 135 Trachootham D, Alexandre J, Huang P . Targeting cancer cells by ROS-mediated mechanisms: a radical therapeutic approach? Nat Rev Drug Discov 2009; 8: 579–591. 136 P elicano H, Feng L, Zhou Y, Carew JS, Hileman EO, P lunkett W et al. Inhibition of mitochondrial respiration: a novel strategy to enhance drug-induced apoptosis in human leukemia cells by a reactive oxygen species-mediated mechanism. J Biol Chem 2003; 278: 37832–37839. 137 Ren YR, P an F, P arvez S, Fleig A, Chong CR, Xu J et al. Clofazimine inhibits human Kv1.3 potassium channel by perturbing calcium oscillation in T lymphocytes. P LoS One 2008; 3: e4009. 138 Felipe A, Bielanska J, Comes N, Vallejo A, Roig S, Ramo ´ n Y, Cajal S et al. Targeting the voltage-dependent K( þ ) channels Kv1.3 and Kv1.5 as tumor biomarkers for cancer detection and prevention. Curr Med Chem 2012; 19: 661– 674. 139 Leanza L, O’ Reilly P , Doyle A, Venturini E, Zoratti M, Szegezdi E et al. Correlation between potassium channel expression and sensitivity to drug-induced cell death in tumor cell lines. Curr P harm Des 2013; P MID:23701546. 140 Liu J, Mu C, Yue W, Li J, Ma B, Zhao L et al. A diterpenoid derivate compound targets selenocysteine of Thioredoxin Reductases and induces Bax/Bak-inde- pendent apoptosis. Free Radic Biol Med 2013; 63: 485–494. 141 Ni B, Ma Q, Li B, Zhao L, Liu Y, Zhu Y, Chen Q. P henylarsine oxide induces apoptosis in Bax- and Bak-deficient cells through upregulation of Bim. Clin Cancer Res 2012; 18: 140–145. 142 Lei X, Chen Y, Du G, Yu W, Wang X, Qu H et al. Gossypol induces Bax/Bak- independent activation of apoptosis and cytochrome c release via a con- formational change in Bcl-2. FASEB J 2006; 20: 2147–2149. 143 LeBlanc H, Lawrence D, Varfolomeev E, Totpal K, Morlan J, Schow P et al. Tumor-cell resistance to death receptor-induced apoptosis through mutational inactivation of the proapoptotic Bcl-2 homolog Bax. Nat Med 2002; 8: 274–281. 144 McCurrach ME, Connor TM, Knudson CM, Korsmeyer SJ, Lowe SW. Bax-deficiency promotes drug resistance and oncogenic transformation by attenuating p53-dependent apoptosis. P roc Natl Acad Sci USA 1997; 94: 2345–2349. 145 Meijerink JP , Mensink EJ, Wang K, Sedlak TW, Slo ¨ etjes AW, de Witte T et al. Hematopoietic malignancies demonstrate loss-of-function mutations of BAX. Blood 1998; 91: 2991–2997. 146 Ionov Y, Yamamoto H, Krajewski S, Reed JC, P erucho M. Mutational inactivation of the proapoptotic gene BAX confers selective advantage during tumor clonal evolution. P roc Natl Acad Sci USA 2000; 97: 10872–10877. 147 Sassi N, De Marchi U, Fioretti B, Biasutto L, Gulbins E, Francolini F et al. An investigation of the occurrence and properties of the mitochondrial intermediate conductance Ca 2 þ -activated K þ channel mtKCa3.1. Biochim Biophys Acta Bioenergetics 2010; 1797: 1260–1267. 148 Stowe DF, Gadicherla AK, Zhou Y, Aldakkak M, Cheng Q, Kwok WM et al. P rotection against cardiac injury by small Ca( 2 þ )-sensitive K( þ ) channels identified in guinea pig cardiac inner mitochondrial membrane. Biochim Biophys Acta 2013; 1828: 427–442 2013. 149 Dolga AM, Netter MF, P erocchi F, Doti N, Meissner L, Tobaben S et al. Mitochondrial small conductance SK2 channels prevent glutamate-induced oxytosis and mitochondrial dysfunction. J Biol Chem 2013; 288: 10792–10804. 150 Singh H, Stefani E, Toro L. Intracellular BK(Ca) (iBK(Ca)) channels. J P hysiol 2012; 23: 5937–5947. 151 DebskaVielhaber G, Godlewski MM, Kicinska A, Skalska J, Kulawiak B, P iwonska M, Zablocki K, Kunz WS, Szewczyk A, Motyl T. Large-conductance K þ channel openers induce death of human glioma cells. J P hysiol P harmacol 2009; 60: 27–36. 152 Wrzosek A, Tomaskova Z, Ondrias K, Lukasiak A, Szewczyk A. The potassium channel opener CGS7184 activates Ca 2 þ release from the endoplasmic reticu- lum. Eur J P harmacol 2012; 690: 60–67. 153 Ruszna ´k Z, Bakondi G, Kosztka L, P ocsai K, Dienes B, Fodor J et al. Mitochondrial expression of the two-pore domain TASK-3 channels in malignantly transformed and non-malignant human cells. Virchows Arch 2008; 452: 415–426. 154 Kova ´cs I, P ocsai K, Czifra G, Sarkadi L, Szu ¨ cs G, Nemes Z et al. TASK-3 immunoreactivity is present but shows differential distribution in the human gastro- intestinal tract. Virchows Arch 2005; 446: 402–410. 155 Kosztka L, Ruszna ´k Z, Nagy D, Nagy Z, Fodor J, Szucs G et al. Inhibition of TASK-3 (KCNK9) channel biosynthesis changes cell morphology and decreases both DNA content and mitochondrial function of melanoma cells maintained in cell culture. Melanoma Res 2011; 21: 308–322. 156 Toczy"owska-Mamin ´ ska R, Olszewska A, Laskowski M, Bednarczyk P , Skowronek K, Szewczyk A. P otassium Channel in the Mitochondria of Human Keratinocytes. J Invest Dermatol 2013; doi:10.1038/jid.2013.422. 157 Lee GW, P ark HS, Kim EJ, Cho YW, Kim GT, Mun YJ et al. Reduction of breast cancer cell migration via up-regulation of TASK-3 two-pore domain K þ channel. Acta P hysiol 2012; 204: 513–524. 158 Innamaa A, Jackson L, Asher V, Van Shalkwyk G, Warren A, Hay D et al. Expression and prognostic significance of the oncogenic K2P potassium channel KCNK9 (TASK-3) in ovarian carcinoma. Anticancer Res 2013; 33: 1401–1408. 159 Fedorenko A, Lishko P V, Kirichok Y. Mechanism of fatty-acid-dependent UCP 1 uncoupling in brown fat mitochondria. Cell 2012; 151: 400–413. 160 Hoang T, Smith MD, Jelokhani-Niaraki M. Toward understanding the mechanism of ion transport activity of neuronal uncoupling proteins UCP 2, UCP 4, and UCP 5. Biochemistry 2012; 51: 4004–4014. 161 Cannon B, Nedergaard J. Brown adipose tissue: function and physiological sig- nificance. P hysiol Rev 2004; 84: 277–359. 162 Baffy G. Uncoupling protein-2 and cancer. Mitochondrion 2010; 10: 243–252. 163 Baffy G, Derdak Z, Robson SC. Mitochondrial recoupling: a novel therapeutic strategy for cancer? Br J Cancer 2011; 105: 469–474. 164 Cannon B, Shabalina IG, Kramarova TV, P etrovic N, Nedergaard J. Uncoupling proteins: a role in protection against reactive oxygen species--or not? Biochim Biophys Acta 2006; 1757: 449–458. 165 Arsenijevic D, Onuma H, P ecqueur C, Raimbault S, Manning BS, Miroux B et al. Disruption of the uncoupling protein-2 gene in mice reveals a role in immunity and reactive oxygen species production. Nat Genet 2000; 26: 435–439. 166 Derdak Z, Mark NM, Beldi G, Robson SC, Wands JR, Baffy G. The mitochondrial uncoupling protein-2 promotes chemoresistance in cancer cells. Cancer Res 2008; 68: 2813–2819. 167 Samudio I, Fiegl M, McQueen T, Clise-Dwyer K, Andreeff M. The Warburg effect in leukemia-stroma cocultures is mediated by mitochondrial uncoupling associated with uncoupling protein 2 activation. Cancer Res 2008; 68: 5198–5205. 168 Ayyasamy V, Owens KM, Desouki MM, Liang P , Bakin A, Thangaraj K et al. Cellular model of Warburg effect identifies tumor promoting function of UCP 2 in breast cancer and its suppression by genipin. P LoS One 2011; 6: e24792. 169 Sanchez-Alvarez R, Martinez-Outschoorn UE, Lamb R, Hulit J, Howell A, Gandara R et al. Mitochondrial dysfunction in breast cancer cells prevents tumor growth: understanding chemoprevention with metformin. Cell Cycle 2013; 12: 172– 182. 170 De Stefani D, Raffaello A, Teardo E, Szabo ` I, Rizzuto R. A forty-kilodalton protein of the inner membrane is the mitochondrial calcium uniporter. Nature 2011; 476: 336–340. 171 Baughman JM, P erocchi F, Girgis HS, P lovanich M, Belcher-Timme CA, Sancak Y et al. Integrative genomics identifies MCU as an essential component of the mitochondrial calcium uniporter. Nature 2011; 476: 341–345. 172 P inton P , Ferrari D, Rapizzi E, Di Virgilio F, P ozzan T, Rizzuto R. The Ca 2 þ con- centration of the endoplasmic reticulum is a key determinant of ceramideinduced apoptosis: significance for the molecular mechanism of Bcl-2 action. EMBO J 2001; 20: 2690–2701. 173 Mallilankaraman K, Doonan P , Ca ´rdenas C, Chandramoorthy HC, Mu ¨ ller M, Miller R et al. MICU1 is an essential gatekeeper for MCU-mediated mitochondrial Ca( 2 þ ) uptake that regulates cell survival. Cell 2012; 151: 630–644. 174 Raffaello A, De Stefani D, Sabbadin D, Teardo E, Merli G, P icard A et al. The mitochondrial calcium uniporter is a multimer that can include a dominant- negative pore-forming subunit. EMBO J 2013; 32: 2362–2376. 175 Marchi S, Lupini L, P atergnani S, Rimessi A, Missiroli S, Bonora M et al. Downregulation of the Mitochondrial calcium uniporter by cancer-related miR-25. Curr Biol 2013; 23: 58–63. 176 Curry MC, P eters AA, Kenny P A, Roberts-Thomson SJ, Monteith GR. Mitochondrial calcium uniporter silencing potentiates caspase-independent cell death in MDA-MB-231 breast cancer cells. Biochem Biophys Res Commun 2013; 434: 695–700. 177 Arvizo RR, Moyano DF, Saha S, Thompson MA, Bhattacharya R, Rotello VM et al. P robing novel roles of the mitochondrial uniporter in ovarian cancer cells using nanoparticles. J Biol Chem 2013; 288: 17610–17618. 178 Davis FM, P arsonage MT, Cabot P J, P arat MO, Thompson EW, Roberts-Thomson SJ et al. Assessment of gene expression of intracellular calcium channels, pumps and exchangers with epidermal growth factor-induced epithelial-mesenchymal transition in a breast cancer cell line. Cancer Cell Int 2013; 13: 76. 179 P an X, Liu J, Nguyen T, Liu C, Sun J, Teng Y et al. The physiological role of mitochondrial calcium revealed by mice lacking the mitochondrial calcium uniporter. Nat Cell Biol 2013; 15: 1464–1472. 180 Sancak Y, Markhard AL, Kitami T, Kova ´cs-Bogda ´n E, Kamer KJ, Udeshi ND et al. EMRE is an essential component of the mitochondrial calcium uniporter com- plex. Science 2013; 342: 1379–1382. 181 Chaudhuri D, Sancak Y, Mootha VK, Clapham DE. MCU encodes the pore con- ducting mitochondrial calcium currents. Elife 2013; 2: e00704.

12

182 P lovanich M, Bogorad RL, Sancak Y, Kamer KJ, Strittmatter L, Li AA et al. MICU2, a paralog of MICU1, resides within the mitochondrial uniporter complex to regulate calcium handling. P LoS One 2013; 8: e55785. 183 Wang YZ, Zeng WZ, Xiao X, Huang Y, Song XL, Yu Z et al. Intracellular ASIC1a regulates mitochondrial permeability transition-dependent neuronal death. Cell Death Differ 2013; 20: 1359–1369. 184 Friese MA, Craner MJ, Etzensperger R, Vergo S, Wemmie JA, Welsh MJ et al. Acid- sensing ion channel-1 contributes to axonal degeneration in autoimmune inflammation of the central nervous system. Nat Med 2007; 13: 1483–1489. 185 Ko JH, Ko EA, Gu W, Lim I, Bang H, Zhou T. Expression profiling of ion channel genes predicts clinical outcome in breast cancer. Mol Cancer 2013; 12: 106. 186 Smith RA, Hartley RC, Murphy MP . Mitochondria-targeted small molecule ther- apeutics and probes. Antioxid Redox Signal 2011; 15: 3021–3038. 187 Smith RA, Hartley RC, Cocheme ´ HM, Murphy MP . Mitochondrial pharmacology. Trends P harmacol Sci. 2012; 33: 341–352. 188 Neuzil J, Dong LF, Rohlena J, Truksa J, Ralph SJ. Classification of mitocans, anti-cancer drugs acting on mitochondria. Mitochondrion 2013; 13: 199–208. 189 Galluzzi LMorselli E, Kepp O, Tajeddine N, Kroemer G. Targeting p53 to mitochondria for cancer therapy. Cell Cycle 2008; 7: 1949–1955. 190 Fulda S, Scaffidi C, Susin SA, Krammer P H, Kroemer G, P eter ME, Debatin KM. Activation of mitochondria and release of mitochondrial apoptogenic factors by betulinic acid. J Biol Chem 1998; 273: 33942–33948. 191 Cavalieri E, Rigo A, Bonifacio M, Carcereri de P rati A, Guardalben E, Bergamini C, Fato R, P izzolo G, Suzuki H, Vinante F. P ro-apoptotic activity of a-bisabolol in preclinical models of primary human acute leukemia cells. J Transl Med 2011; 9: 45. 192 Guo XP , Zhang XY, Zhang SD. Clinical trial on the effects of shikonin mixture on later stage lung cancer. Zhong Xi Yi Jie He Za Zhi 1991; 11: 598–599. 193 Schonherr R. Clinical relevance of ion channels for diagnosis and therapy of cancer. J Membr Biol 2005; 205: 175–184. 194 Wulff H, Miller MJ, Hansel W, Grissmer S, Cahalan MD, Chandy KG. Design of a potent and selective inhibitor of the intermediate-conductance Ca2 þ - activated K þ channel, IKCa1: a potential immunosuppressant. P roc Natl Acad Sci USA 2000; 97: 8151–8156.

13